Download The AMP-activated protein kinase pathway – new

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Cytokinesis wikipedia , lookup

Histone acetylation and deacetylation wikipedia , lookup

Proteasome wikipedia , lookup

Magnesium transporter wikipedia , lookup

Biochemical switches in the cell cycle wikipedia , lookup

Hedgehog signaling pathway wikipedia , lookup

Protein moonlighting wikipedia , lookup

Protein (nutrient) wikipedia , lookup

SULF1 wikipedia , lookup

List of types of proteins wikipedia , lookup

G protein–coupled receptor wikipedia , lookup

Tyrosine kinase wikipedia , lookup

Apoptosome wikipedia , lookup

Phosphorylation wikipedia , lookup

Signal transduction wikipedia , lookup

Protein phosphorylation wikipedia , lookup

Mitogen-activated protein kinase wikipedia , lookup

Transcript
Commentary
5479
The AMP-activated protein kinase pathway – new
players upstream and downstream
D. Grahame Hardie
Division of Molecular Physiology, Wellcome Trust Biocentre, University of Dundee, Dow Street, Dundee, DD1 5EH, Scotland, UK
(e-mail: [email protected])
Journal of Cell Science 117, 5479-5487 Published by The Company of Biologists 2004
doi:10.1242/jcs.01540
Summary
The AMP-activated protein kinase (AMPK) cascade is a
sensor of cellular energy status. Whenever the cellular
ATP:ADP ratio falls, owing to a stress that inhibits ATP
production or increases ATP consumption, this is amplified
by adenylate kinase into a much larger increase in the
AMP:ATP ratio. AMP activates the system by binding to
two tandem domains on the γ subunits of AMPK, and this
is antagonized by high concentrations of ATP. AMP
binding causes activation by a sensitive mechanism
involving phosphorylation of AMPK by the tumour
suppressor LKB1. Once activated, AMPK switches on
catabolic pathways that generate ATP while switching off
ATP-consuming processes. As well as acting at the level of
the individual cell, the system also regulates food intake
and energy expenditure at the whole body level, in
particular by mediating the effects of hormones and
cytokines such as leptin, adiponectin and ghrelin. A
particularly interesting downstream target recently
identified is TSC2 (tuberin). The LKB1→AMPK→TSC2
pathway negatively regulates the target of rapamycin
(TOR), and this appears to be responsible for limiting
protein synthesis and cell growth, and protecting against
apoptosis, during cellular stresses such as glucose
starvation.
Introduction
A useful analogy can be drawn between the ubiquitous cellular
nucleotides, ATP and ADP, and the chemicals in an electrical
cell or battery. Living cells normally maintain a high ratio of
ATP to ADP (typically around 10:1), which is many orders of
magnitude away from the equilibrium ratio for ATP hydrolysis
under cellular conditions. This represents a store of energy that
can be used to drive energy-requiring processes. Catabolism
(and photosynthesis when available) ‘charges up the battery’
by converting ADP to ATP, whereas most other cellular
processes require ATP hydrolysis and tend to discharge the
battery (Fig. 1). Most cells manage to maintain their ATP:ADP
ratio within fairly narrow limits, which indicates that the rate
of ATP synthesis exactly matches the rate of ATP consumption.
How do they achieve this remarkable feat? In mammalian cells,
it has recently become clear that the AMP-activated protein
kinase (AMPK) system plays a crucial role.
One might expect that this system would respond to changes
in the cellular ADP:ATP ratio. However, if the reaction
catalysed by adenylate kinase (2ADP ↔ ATP + AMP) is
maintained close to equilibrium (which appears to be the case
in most eukaryotic cells), one can easily show that the
AMP:ATP ratio varies as the square of the ADP:ATP ratio
(Hardie and Hawley, 2001). The former ratio is therefore a
much more sensitive indicator of cellular energy status than the
latter, and is the parameter that is monitored by the AMPK
system (Fig. 1).
The AMPK field is developing very rapidly, and this
Commentary cannot hope to be comprehensive. I therefore
focus on recent developments, particularly the identification of
new players acting upstream and downstream of AMPK.
Readers who are particularly interested in aspects related to
medicine or physiology should consult other recent reviews
(Hardie, 2003; Hardie et al., 2003; Carling, 2004; Hardie,
2004a; Hardie, 2004b).
Key words: AMP-activated protein kinase, LKB1, TSC2
Structure of the AMPK complex
Mammalian AMPKs are heterotrimeric complexes consisting
of a catalytic α subunit and regulatory β and γ subunits (Hardie
et al., 2003), each of which is encoded by two or three distinct
genes (α1, α2; β1, β2; γ1, γ2, γ3) (Fig. 2). All possible
combinations of isoforms appear to be able to form complexes
and, along with splice variants and the use of alternative
promoters, this leads to a diverse array of different AMPK
complexes. Although the mammalian kinase is the only
example that is well characterized at the biochemical level,
genes encoding orthologues of the α, β and γ subunits are
found in all eukaryotic species whose genome sequences have
been determined. This even includes the primitive parasite
Giardia lamblia (Hardie et al., 2003), which lacks
mitochondria and, according to molecular phylogenies, is more
than twice as distant from humans as is Saccharomyces
cerevisiae (Adam, 2000). Thus, the existence of AMPK
orthologues appears to be a fundamental feature of all
eukaryotic cells.
What are the functions of the three subunits? The α1 and α2
subunits contain a conventional kinase domain at the Nterminus (Fig. 2), and a phylogenetic tree of the >500 such
domains encoded in the human genome (Manning et al., 2002)
reveals that these lie on a distinct branch containing eleven
others (Fig. 3) that we now refer to as the AMPK-related kinase
5480
Journal of Cell Science 117 (23)
A α subunits
consumptio
n
ATP
Upstream kinases
P
ATP
AMPK
AMP
Adenylate
kinase
ADP
C
N
Kinase
C a t a b o li s m
B β subunits
N
Fig. 1. Physiological role of AMPK in the cell. Catabolism ‘charges
up the battery’ by converting ADP to ATP (bottom curved arrow)
whereas ATP-consuming processes convert ATP to ADP (top curved
arrow). If a cellular stress causes the rate of catabolism to fail to keep
pace with the rate of ATP consumption, ADP levels will rise and
ATP levels will fall. ADP is converted into AMP by adenylate kinase
and this, combined with the fall in ATP, will activate AMPK. AMPK
then promotes the restoration of energy balance by stimulating
catabolism and inhibiting ATP-consuming processes.
family (Lizcano et al., 2004). The C-terminal regions of the α
subunits are required to form a complex with the β and γ
subunits (Crute et al., 1998). Comparison of β subunit
sequences from different species revealed that they contained
two conserved regions originally termed the KIS and ASC
domains (Jiang and Carlson, 1997). Recent work has shown
that, in mammals, the ASC domain is sufficient for the
formation of a complex with the α and γ subunits, whereas the
KIS domain is not (as previously thought) involved in subunit
interactions but is in fact a glycogen-binding domain (GBD)
(Hudson et al., 2003) (Fig. 2). According to the protein families
(Pfam) database, the GBD is a member of the isoamylase-N
domain family (see entry PF02922 at http://www.sanger.ac.uk/
Software/Pfam/). These non-catalytic domains are usually
found in enzymes that metabolize the α1→6 branch points in
α1→4-linked glucans, such as glycogen and starch. Two lines
of evidence confirmed that the GBD was indeed involved in
glycogen binding. First, when the α, β and γ subunits are coexpressed in cultured human cells with a green fluorescent
protein (GFP) tag on α, fluorescence is associated with
glycogen-containing granules in the cytoplasm, but this is
abolished if the GBD is deleted from the β subunit (Hudson et
al., 2003). Second, the GBD from rat β1 synthesized in bacteria
binds to glycogen in cell-free assays (Polekhina et al., 2003).
The exact function(s) of the GBD remains unclear, although
there are several experimental findings linking AMPK with
glycogen; however, these observations are currently difficult to
synthesize into a single, all-encompassing hypothesis. The
GBD does cause a partial localization of AMPK to glycogen
particles, where one of its known downstream targets –
glycogen synthase – resides (Hudson et al., 2003). There is also
indirect evidence that glycogen regulates AMPK activity: in
both rat (Wojtaszewski et al., 2002) and human (Wojtaszewski
et al., 2003) skeletal muscle, a high content of glycogen
represses activation of AMPK. This makes physiological sense
because if muscle glycogen content is high it tends to be used
preferentially as fuel and, although AMPK activation
stimulates the usage by muscle of alternative fuels such as
blood glucose and fatty acids, it is not required for glycogen
breakdown or glycolysis. As yet, repression of AMPK
Complex
formation
ASC C
KIS
Glycogenbinding
Complex
formation
C γ subunits
N CBS1 CBS2 CBS3 CBS4 C
γ1
CBS1 CBS2 CBS3 CBS4 C
γ2
N
N
CBS1 CBS2 CBS3 CBS4 C
AMP/ATP binding
γ3
Fig. 2. Conserved domains in AMPK subunits: (A) α subunits, (B) β
subunits and (C) γ subunits. The proposed function of each domain is
indicated. The two isoforms of the α (α1, α2) and β (β1, β2)
subunits have very similar structures, but the three isoforms of the γ
subunit (γ1, γ2, γ3) contain variable N-terminal regions of unknown
function, and are drawn separately. [Redrawn from Hardie et al.
(Hardie et al., 2003).]
activation by glycogen has not been reproduced in a cell-free
system; so the molecular mechanism remains unclear. Given
the ample evidence (discussed further below) that the AMPK
system senses the immediate availability of cellular energy in
the form of adenine nucleotides, an attractive hypothesis is that
it can also sense longer-term reserves in the form of glycogen.
Whatever its functions, the GBD is found universally in the β
subunit orthologues of all eukaryotes, which suggests that its
function(s) are ancient and important.
AMPK-α2
AMPK-α1
BRSK2
BRSK1
NuaK1 (Ark5)
NuaK2 (SNARK)
QSK
SIK
QIK (SIK2)
MARK4
MARK1
MARK3 (PAR1A/C-TAK1)
MARK2
Fig. 3. Phylogenetic tree of the AMPK-related protein kinases.
[Based on Manning et al. (Manning et al., 2002).]
AMP-activated protein kinase
Following their variable N-terminal regions, the γ subunits
of AMPKs contain four tandem repeats of a sequence motif
first identified by Bateman as a ‘CBS domain’ (Bateman, 1997)
(Fig. 2). These motifs contain about 60 residues and are found
in various proteins in archaea, bacteria and eukaryotes: the
acronym derives from the enzyme cystathionine β-synthase,
which has a pair of such motifs at the C-terminus. They
invariably occur as tandem pairs, and the crystal structure of a
bacterial inosine-5′-monophosphate (IMP) dehydrogenase
(Zhang et al., 1999) shows that the repeats in a pair associate
closely together through hydrophobic interactions. As it is now
clear that the basic functional unit is a dimer, Kemp has
suggested the term ‘Bateman domain’ to refer to the structure
formed by two tandem repeats (Kemp, 2004). Initially, the
function(s) of Bateman domains was not known, but recent
studies have revealed that Bateman domains in the γ subunits
represent the regulatory AMP- and ATP-binding sites of the
AMPK complex (Scott et al., 2004). The N-terminal and Cterminal Bateman domains from γ2 can each bind one molecule
of AMP or ATP in a mutually exclusive manner, whereas a
construct containing both Bateman domains binds two
molecules of AMP or ATP with strong positive cooperativity.
The finding that there are two binding sites for AMP and ATP
in a single αβγ complex had not been anticipated. Strong
evidence that the Bateman domains are indeed the regulatory
AMP- and ATP-binding sites came from studies of mutations
in γ2 that cause a human heart condition, Wolff-ParkinsonWhite syndrome. All five mutations caused both defective
binding of AMP to the isolated Bateman domains (Scott et al.,
2004) and defective activation of the intact αβγ complex by
AMP (Daniel and Carling, 2002; Scott et al., 2004), and the
order of severity of the individual mutations was the same for
both parameters.
Bateman domains from other proteins bind other adenosine
derivatives (Scott et al., 2004). Those of IMP dehydrogenase
and CLC chloride channels bind ATP, whereas those of
cystathionine β-synthase bind S-adenosyl methionine, the
common feature being the binding of an adenosine-containing
ligand that activates protein function. Intriguingly, in all of
these proteins, point mutations that interfere with ligand
binding cause a variety of human hereditary diseases ranging
from retinitis pigmentosa to epilepsy (Scott et al., 2004).
Regulation of the AMPK complex
AMPK is activated by 5′-AMP in three distinct ways, all of
which are antagonized by high concentrations of ATP.
First, AMP binding causes allosteric activation of
mammalian AMPK. The magnitude of this effect varies
between AMPK isoforms but is typically fivefold or less. A
puzzling finding is that, although the γ subunits of the yeast
and plant orthologues contain two Bateman domains that are
very similar to those of the mammalian enzyme, allosteric
activation of the yeast and plant kinases has not been
demonstrated. This might simply be due to technical problems:
because adenylate kinase converts ADP formed in the kinase
assay to AMP, one cannot demonstrate AMP activation unless
adenylate kinase is completely removed. AMP does activate
the AMPK orthologue from Drosophila (Pan and Hardie,
2002), which shows that this regulatory mechanism is at least
conserved between mammals and insects.
5481
Second, AMP binding makes AMPK a better substrate for
the upstream kinase. The upstream kinase, recently identified
as the tumour suppressor LKB1 (see below), activates the
kinase by phosphorylation of the α subunit at a specific
threonine residue (Thr172) (Hawley et al., 1996) in the socalled activation loop, within which phosphorylation causes
activation of many other protein kinases. As phosphorylation
of Thr172 causes at least 50- to 100-fold activation, it is
quantitatively much more important than the allosteric
activation. Like the allosteric effect, activation by the upstream
kinase is stimulated by AMP and inhibited by high
concentrations of ATP (Hawley et al., 1996). AMP stimulation
is only observed if the intact αβγ complex, and not the isolated
kinase domain (which lacks the AMP-binding sites), is used as
the substrate (Hawley et al., 2003). This shows that the effect
is due to binding of AMP to the substrate, AMPK, and not to
the upstream kinase. The critical threonine residue, and the
sequence around it, is conserved in the α subunit sequences of
AMPK orthologues from all eukaryotes, and mutation of the
equivalent threonine residue in the yeast α subunit (Snf1p)
causes a total loss of function in vivo (Estruch et al., 1992).
Thus, although it is not yet clear whether the allosteric
activation by AMP is conserved in all eukaryotes, the
regulation by phosphorylation is certainly conserved.
Third, AMP binding to AMPK also inhibits
dephosphorylation of Thr172 by protein phosphatases (Davies
et al., 1995). Unlike the allosteric activation, this mechanism
of regulation by AMP has also been reported to operate in
higher plants (Sugden et al., 1999).
Why should AMP activate the system by three mechanisms,
all of which probably involve binding to the same sites on the
γ subunit? Computer simulations have revealed that this causes
the system to be ultrasensitive; thus, a small change in the
initial input (AMP) can produce a much larger relative change
in the final output (Hardie et al., 1999). This is an example of
multistep sensitivity, which arises when a signal molecule
affects more than one step in a cascade (Chock and Stadtman,
1977). Sensitivity in the AMPK system also arises because the
upstream kinase has a very low Km for AMPK (Hardie et al.,
1999), a phenomenon referred to as zero-order ultrasensitivity
(Goldbeter and Koshland, 1981). A factor that might increase
the sensitivity of the AMPK system even further is the recent
finding that the two Bateman domains on the γ subunits bind
AMP cooperatively (see above). Overall, this exquisite
sensitivity might be very important in the ability of AMPK to
maintain the energy state of the cell within narrow limits.
Mammalian AMPK is activated by stresses that
cause ATP depletion
AMPK is therefore activated by any stress that depletes cellular
ATP, such as metabolic poisoning, oxidative stress, hypoxia, or
nutrient deprivation (Table 1). These can be regarded as
pathological stresses that interfere with ATP production. A
physiological stress that activates AMPK by increasing ATP
consumption is exercise in skeletal muscle (Winder and
Hardie, 1996). Indeed, numerous studies now suggest that
AMPK is an important mediator of the varied metabolic
responses to exercise. For full coverage of this topic, the reader
should refer to other reviews (Hardie, 2003; Hardie, 2004a;
Hardie, 2004b). It is becoming clear that AMPK may be
5482
Journal of Cell Science 117 (23)
Table 1. Stresses that activate AMPK
Type of treatment
Treatment
Primary reference
Respiratory chain inhibitors
Antimycin A
Azide
Nitric oxide
Oligomycin
Dinitrophenol
Over-expression of UCP1
Over-expression of UCP3
Arsenite
Heat shock
Oxidative stress
Osmotic stress
Ischaemia
Hypoxia
Low glucose
Exercise (muscle)
Contraction (muscle)
Metformin
Phenformin
Rosiglitazone
Pioglitazone
Witters et al., 1991
Witters et al., 1991
Almeida et al., 2004
Marsin et al., 2000
Witters et al., 1991
Matejkova et al., 2004
Schrauwen et al., 2004
Corton et al., 1994
Corton et al., 1994
Choi et al., 2001
Fryer et al., 2002
Kudo et al., 1995
Marsin et al., 2000
Salt et al., 1998
Winder and Hardie, 1996
Hutber et al., 1997; Vavvas et al., 1997
Zhou et al., 2001
Hawley et al., 2003
Fryer et al., 2002
Saha et al., 2004
ATP synthase inhibitor
Mitochondrial uncouplers
TCA cycle inhibitor
Environmental stresses
Metabolic stresses
Biguanide drugs
Thiazolidinedione drugs
responsible for many of the health benefits of regular exercise,
particularly with respect to prevention and treatment of obesity
and Type 2 diabetes. Medical interest in the AMPK system was
further stimulated by the finding that it is activated by two of
the main classes of drugs currently used for treatment of Type
2 diabetes: biguanides and thiazolidinediones (Table 1). Both
classes of drug inhibit complex I of the respiratory chain
(Owen et al., 2000; Brunmair et al., 2004) and this might be
how they activate AMPK, although in the case of metformin it
has proved difficult to measure changes in cellular AMP:ATP
ratios (Fryer et al., 2002; Hawley et al., 2002). With the
possible exception of osmotic stress (Fryer et al., 2002), all of
the stresses that activate AMPK are associated with changes in
the cellular AMP:ATP ratio.
Physiological roles of AMPK orthologues in yeast
and plants
The SNF1 (for ‘sucrose, non-fermenting’) and SNF4 genes
respectively encode the α and γ subunits of the yeast SNF1
complex, and were originally identified by genetic screens for
mutations that prevent growth on sucrose and non-fermentable
carbon sources (Hardie et al., 1998). Growth on sucrose
requires the expression of the enzyme that breaks it down
(invertase), whereas growth on non-fermentable carbon
sources requires expression of mitochondrial genes needed for
oxidative metabolism. Expression of all of these genes is
repressed by glucose, and the SNF1 and SNF4 genes are
required for their derepression. One mechanism by which this
is mediated is the phosphorylation of the repressor protein
Mig1p by the SNF1 complex (Smith et al., 1999).
Phosphorylation causes Mig1p to bind to a nuclear export
protein that promotes its removal from the nucleus (DeVit and
Johnston, 1999).
Thus, the primary role of the AMPK orthologue in yeast is
in the response to glucose starvation. In cultured cells from
mammals (Salt et al., 1998) and Drosophila (Pan and Hardie,
2002), AMPK is also activated by glucose deprivation; so this
seems to have been an ancient role for the system. In green
plants, darkness is the equivalent of starvation: plants
synthesize carbohydrates by photosynthesis during light
periods and then rely on breakdown of carbohydrate reserves
(in the form of starch) during dark periods. The moss
Physcomitrella patens has two genes encoding orthologues of
the AMPK α subunit. Disruption of both genes produces a
complete failure to grow unless the plants are maintained under
constant illumination (Thelander et al., 2004). Thus, the
response to starvation also seems to be a fundamental role of
AMPK orthologues in green plants.
Regulation of whole body energy intake and
expenditure
Although AMPK orthologues are present in single-celled
eukaryotes, and AMPK initially appeared to regulate energy
balance in a cell-autonomous manner, it has recently become
clear that in multicellular organisms it also plays the same role
at the whole body level. The first indications of this came with
findings that AMPK mediates stimulation of fatty oxidation in
skeletal muscle by leptin (Minokoshi et al., 2002) and
adiponectin (Tomas et al., 2002; Yamauchi et al., 2002). These
two hormones are so-called adipokines secreted by adipocytes,
and are thought to represent signals that body lipid stores are
adequate. AMPK activation in muscle also increases glucose
uptake (Merrill et al., 1997) and mitochondrial biogenesis
(Zong et al., 2002), and leptin and adiponectin would therefore
increase energy expenditure by stimulating oxidation of
glucose as well as fatty acids in muscle. As well as its effects
on muscle, high leptin levels in vivo can cause adipocytes to
switch from net synthesis to net oxidation of fatty acids, and
this is associated with AMPK activation in the cells (Orci et
al., 2004). Thus, AMPK stimulates ‘burning’ of fat when its
storage is excessive (although why this fails in obese
individuals remains unclear).
Remarkably, as well as increasing energy expenditure by
stimulating fatty acid and glucose oxidation, AMPK also
appears to regulate energy intake in mammals. In the regions
of the hypothalamus involved in regulation of satiety and food
intake, AMPK is inhibited in rodents by high levels of glucose,
leptin and insulin, all of which repress food intake, whereas it
AMP-activated protein kinase
is activated by ghrelin, a gut hormone that stimulates food
intake (Andersson et al., 2004; Minokoshi et al., 2004). These
changes in AMPK activity seem to trigger the subsequent
changes in food intake, because direct injection of the AMPK
activator 5-aminoimidazole 4-carboxamide (AICA) riboside
into the hypothalamus increases food intake (Andersson et al.,
2004), whereas expression of activated and dominant-negative
inhibitory mutants of AMPK in the hypothalamus stimulates
and inhibits food intake, respectively (Minokoshi et al., 2004).
One interesting aspect of these findings is that leptin inhibits
AMPK in the hypothalamus yet activates it in skeletal muscle,
raising interesting questions about the mechanisms involved. It
is also intriguing that the AMPK system appears to be involved
in increasing food intake in response to lowered blood glucose
levels in mammals, just as its orthologue in yeast, the SNF1
complex, is involved in the response to lowered glucose levels
in the extracellular medium.
Identification of the upstream kinase, LKB1-STRADMO25
Although it had been clear for many years that the mechanism
of AMPK activation by stresses involves an upstream kinase,
which was partially purified from rat liver (Hawley et al.,
1996), its identity remained a mystery. However, recently, we
(Sutherland et al., 2003) and others (Hong et al., 2003a)
identified three closely related protein kinases (Elm1p, Tak1p,
Tos3p) capable of acting upstream of the yeast SNF1 complex
in vivo. One of the closest sequence matches to these in
humans was the protein kinase LKB1, and compelling
evidence now indicates that this is the key upstream kinase in
mammalian cells (Hawley et al., 2003; Woods et al., 2003;
Shaw et al., 2004). LKB1 forms a complex with two accessory
subunits: STRAD and MO25 (Baas et al., 2003; Boudeau et
al., 2003). The upstream kinase purified from rat liver is a
complex between LKB1, STRAD and MO25, and all three
subunits are required for full activity (Hawley et al., 2003).
LKB1 was first identified as the gene mutated in human
Peutz-Jeghers syndrome (PJS) (Hemminki et al., 1998; Jenne
et al., 1998). Subjects with PJS are heterozygous for a loss-offunction mutation in the LKB1 gene. They have unusual skin
pigmentation and also suffer from numerous gastrointestinal
polyps that are classed as hamartomas, benign tumours caused
by abnormal cell growth, in which the normal differentiated
state of the cells is maintained. Subjects with PJS also have an
increased risk of developing malignant tumours, which led to
LKB1 being designated as a tumour suppressor.
Heterozygous mice that lack one functional copy of the
LKB1 gene develop extensive gastrointestinal polyps, as in
the human syndrome; however, there are conflicting reports
as to whether these are caused by loss of heterozygosity, i.e.
loss of the remaining functional copy of the gene (Bardeesy
et al., 2002; Miyoshi et al., 2002; Rossi et al., 2002). One
strain of these mice has increased incidence of liver cancers,
which is associated with loss of heterozygosity (Nakau et al.,
2002). Homozygous LKB1-knockout mice die during
embryonic development, but immortalized mouse embryo
fibroblasts (MEFs) can be derived from the embryos, and
have provided an important tool that proved that LKB1 is
necessary for AMPK activation (Hawley et al., 2003; Shaw
et al., 2004).
5483
Are the tumour suppressor effects of LKB1 due to its ability
to activate AMPK? This is certainly possible, because AMPK
activation in HepG2 cells causes a cell-cycle arrest in G1 phase
through stabilization of p53 (Imamura et al., 2001), and
expression of LKB1 in HeLa or G361 cells (tumour cell lines
that do not normally express the kinase) also causes a p53dependent G1 arrest (Tiainen et al., 1999; Tiainen et al., 2002).
However, LKB1 lies upstream of all of the AMPK-related
kinases shown in Fig. 3 (Lizcano et al., 2004). Although the
four MARKs appear to have functions in regulating cell
polarity (Baas et al., 2004), the functions and downstream
targets of the other AMPK-related kinases are not well
understood. Thus, some of the tumour suppressor effects of
LKB1 might be mediated by one or more of the AMPK-related
kinases rather than AMPK itself.
Is the LKB1 complex itself regulated? Although this
possibility cannot be excluded, recent results suggest that it is
constitutively active. First, during stimulation of cultured
muscle cells by several stress treatments that activate AMPK,
LKB1 activity was constant (Woods et al., 2003). Second,
during activation of AMPK in MEFs by the anti-diabetic drug
phenformin, there was no change in LKB1 activity (Lizcano et
al., 2004), despite the fact that activation requires the presence
of LKB1 (Hawley et al., 2003). Third, during treatment of rat
skeletal muscle with different AMPK-activating stimuli, the
activities of LKB1, and several of the AMPK-related kinases
downstream of LKB1, did not change (Sakamoto et al., 2004).
Downstream targets of AMPK
A full discussion of the downstream targets and processes
regulated by AMPK can be found elsewhere (Hardie, 2003;
Hardie et al., 2003; Hardie, 2004a; Hardie, 2004b). Fig. 4
summarizes some of the well-established downstream targets,
and a few of the more recently identified targets are discussed
below. In general, activation of AMPK downregulates
biosynthetic pathways such as fatty acid and cholesterol
biosynthesis, yet switches on catabolic pathways that generate
ATP, such as fatty acid oxidation, glucose uptake and
glycolysis (Fig. 1). It achieves this not only through direct
phosphorylation of metabolic enzymes, but also through
effects on gene expression, such as upregulation of the glucose
transporter GLUT4 and mitochondrial genes in muscle, and
downregulation of genes encoding enzymes of fatty acid
synthesis and gluconeogenesis in liver. Although in no single
case do we completely understand the detailed mechanisms by
which AMPK activation regulates the expression of a particular
gene, it has many effects on individual transcription factors and
coactivators. For example, it upregulates the expression of the
co-activator PGC1α (Terada et al., 2002), which may be
involved in the increased expression of mitochondrial genes
in muscle (Zong et al., 2002), yet it downregulates the
transcription factors SREBP-1c (Zhou et al., 2001) and HNF4α
(Leclerc et al., 2001; Hong et al., 2003b), which may be
involved in decreased expression of genes involved in
lipogenesis, glucose uptake and glycolysis in liver. In muscle,
it causes increased DNA binding by the transcription factors
NRF1 (Bergeron et al., 2001) and MEF2 (Zheng et al., 2001),
which may be involved in regulation of mitochondrial genes
and GLUT4, respectively. Finally, it phosphorylates the
ubiquitous co-activator p300 at a specific site (Ser89), reducing
5484
Journal of Cell Science 117 (23)
Fatty acid
synthesis
Gluconeogenesis
Mitochondrial
biogenesis
GLUT4
Fatty acid
oxidation
FAS, ACC1
PEPCK
G6Pase
PGC1α
MEF2
NRF1
Protein
synthesis
Fatty acid
synthesis
ACC1
?
SREBP1c
HNF4α
Cholesterol
(isoprenoid)
synthesis
ACC2
?
HMGR
EF2
?
AMPK
Cell growth
& protein
synthesis
Glycogen
synthesis
GS
TSC2
TOR
HSL
CFTR
?
?
PFK2
eNOS
nNOS
Cl–/fluid
secretion
Lipolysis
GLUT1
GLUT4
CD36
FAT
Glycolysis
Blood
flow
Fatty acid
uptake
Glucose
uptake
Fig. 4. Targets for AMPK. Target proteins and processes activated by AMPK activation are shown in green, and those inhibited by AMPK
activation are shown in red. Where the effect is caused by a change in gene expression, an upward-pointing green arrow next to the protein
indicates an increase, whereas a downward-pointing red arrow indicates a decrease in expression. Abbreviations: ACC1/ACC2, 1 (α) and 2 (β)
isoforms of acetyl-CoA carboxylase; CD36/FAT, CD36/fatty acid translocase; CFTR, cystic fibrosis transmembrane regulator; EF2, elongation
factor-2; eNOS/nNOS. endothelial/neuronal isoforms of nitric oxide synthase; FAS, fatty acid synthase; G6Pase, glucose-6-phosphatase;
GLUT1/4, glucose transporters; GS, glycogen synthase; HMGR, 3-hydroxy-3-methyl-CoA reductase; HSL, hormone-sensitive lipase; MEF2,
myocyte-specific enhancer factor-2; NRF1, nuclear respiratory factor-1; PEPCK, phosphoenolpyruvate carboxykinase; PGC1α, peroxisome
proliferator-activated receptor-γ co-activator-1α; TOR, mammalian target of rapamycin.
its ability to bind to nuclear hormone receptors and thus
activate their target genes (Yang et al., 2001).
Another mechanism by which AMPK activation affects the
expression of proteins is through effects on mRNA stability.
Through an undefined mechanism, AMPK activation reduces
the cytoplasmic level of the RNA-binding protein HuR, which
stabilizes specific mRNAs in the cytoplasm by binding to their
3′-untranslated regions (Wang et al., 2002). Target mRNAs for
HuR include proteins that regulate the cell cycle, such as cyclin
A, cyclin B1 and p21. The same group showed that activation
of AMPK by an elevated AMP:ATP ratio in senescent
fibroblasts in culture can explain many of the features of the
senescent phenotype (Wang et al., 2003). Indeed, LKB1–/– MEF
cells, in which AMPK cannot be activated by agents that elevate
cellular AMP levels (phenformin) or that mimic the effects of
AMP (AICA riboside) (Hawley et al., 2003), are resistant to
passage-induced senescence (Bardeesy et al., 2002).
Another key biosynthetic pathway that is downregulated by
AMPK is translation, which can account for around 20% of
energy turnover in growing cells and is particularly sensitive
to decreases in ATP synthesis (Buttgereit and Brand, 1995).
Inhibition occurs by at least two mechanisms: (1)
phosphorylation and activation of elongation factor-2 kinase
(Horman et al., 2002; Browne et al., 2004), which switches off
the elongation step in translation; and (2) inhibition of the
target of rapamycin (TOR) pathway (Bolster et al., 2002;
Krause et al., 2002; Kimura et al., 2003), which is a major
positive stimulus for protein synthesis, cell growth and cell
size. The TOR pathway is activated by growth factors and
amino acids, and is thought to stimulate translation, and hence
cell growth, by two mechanisms: (1) activation of ribosomal
protein S6 kinase (S6K1); and (2) increased phosphorylation
of elongation factor-4E binding protein 1 (4E-BP1), which
stimulates the initiation step of translation (Carrera, 2004).
Recent studies suggest that inhibition of the TOR pathway by
AMPK might occur through phosphorylation of TSC2 (Inoki
et al., 2003) (Fig. 5). TSC1 and TSC2 (also known as hamartin
and tuberin) are partners that form a stable complex in the cell,
and mutations in either leads to the human disease tuberous
sclerosis complex. This is similar to PJS in that it is an
AMP-activated protein kinase
Growth
factors
Stress
AMP
LKB1
PtdIns(3,4,5)P3
AMPK
PKB/Akt
TSC1-TSC2
PDK1
Amino
acids
TOR
S6K1
4E-BP1
Protein synthesis
Cell growth
Fig. 5. Regulation of protein synthesis and cell growth by AMPK
and PKB/Akt by the mTOR pathway. Cellular stresses activate
AMPK because the increase in AMP promotes its phosphorylation
by LKB1; whereas growth factors activate PKB/Akt because the
increase in phosphatidylinositol (3,4,5)-trisphosphate
[PtdIns(3,4,5)P3] promotes its phosphorylation by PDK1. AMPK
and PKB/Akt phosphorylate TSC2 at different sites, and this
stimulates or inhibits, respectively, the ability of the TSC1-TSC2
complex to inhibit TOR. Amino acids also stimulate TOR through
the TSC complex. TOR in turn stimulates protein synthesis, and
hence cell growth, through ribosomal protein S6 kinase 1 (S6K1) and
elongation factor-4E binding protein 1 (4E-BP1). The molecular
events immediately upstream and downstream of TOR in this
pathway are not shown in detail and remain incompletely
understood.
autosomal dominant condition resulting in hamartomas,
although they are not restricted to the intestine. Studies in
Drosophila suggest that the TSC1-TSC2 complex negatively
regulates cell growth, and acts upstream of TOR to cause its
inhibition (Gao et al., 2002). AMPK phosphorylates TSC2 at
two sites; by mutating these site, Guan and colleagues provided
evidence that these phosphorylation events increase the ability
of the TSC complex to inhibit the TOR pathway, and that
phosphorylation of TSC2 by AMPK is involved in the
mechanism by which the TSC complex inhibits cell growth and
protects against apoptosis during glucose starvation (Inoki et
al., 2003). Interestingly, Shaw et al. reported that LKB1–/– MEF
cells are more susceptible to apoptosis in response to AMPKactivating stresses (Shaw et al., 2004). Given that the
LKB1→AMPK→TSC2 pathway (Fig. 5) seems to limit cell
growth and cell size by inhibiting TOR, the idea that the
hamartomas caused by heterozygosity for mutations in LKB1
(which lies upstream of AMPK) are related to those caused by
heterozygosity for mutations in TSC2 (which lies downstream
of AMPK) is attractive, although why the lesions are restricted
to the intestine in the case of LKB1 remains unclear. However,
it is also possible that the lesions in PJS are caused in part by
dysregulation of downstream kinases other than AMPK, such
as the MARKs.
5485
Conclusions and perspectives
The identification of LKB1 as the upstream kinase ended one
era in research on AMPK but initiated another. In particular,
the work on TSC2 raises the intriguing possibility, as discussed
previously (Shaw et al., 2004), that the AMPK system acts as
an ‘energy checkpoint’ that determines whether cellular energy
status is sufficient before the cell commits to a programme of
growth triggered by the TOR pathway. There are also
indications that AMPK activation in response to low energy
status inhibits cell division through effects on DNA replication
and mitosis, although the detailed mechanisms remain to be
elucidated. Finally, the AMPK system seems to have evolved
in single-celled eukaryotes, where it might have had a primary
role in orchestrating the response to glucose starvation.
However, the recent findings that AMPK is regulated by
cytokines such as leptin and adiponectin in mammals show
that, during the evolution of multicellular organisms, it became
involved in the regulation of energy intake and expenditure at
the whole body level.
Studies in the author’s laboratory are supported by the Wellcome
Trust and by a contract (QLG1-CT-2001-01488) from the European
Commission.
References
Adam, R. D. (2000). The Giardia lamblia genome. Int. J. Parasitol. 30, 475484.
Almeida, A., Moncada, S. and Bolanos, J. P. (2004). Nitric oxide switches
on glycolysis through the AMP protein kinase and 6-phosphofructo-2kinase pathway. Nat. Cell Biol. 6, 45-51.
Andersson, U., Filipsson, K., Abbott, C. R., Woods, A., Smith, K.,
Bloom, S. R., Carling, D. and Small, C. J. (2004). AMP-activated protein
kinase plays a role in the control of food intake. J. Biol. Chem. 279, 1200512008.
Baas, A. F., Boudeau, J., Sapkota, G. P., Smit, L., Medema, R., Morrice,
N. A., Alessi, D. R. and Clevers, H. C. (2003). Activation of the tumour
suppressor kinase LKB1 by the STE20-like pseudokinase STRAD. EMBO
J. 22, 3062-3072.
Baas, A. F., Kuipers, J., van der Wel, N. N., Batlle, E., Koerten, H. K.,
Peters, P. J. and Clevers, H. C. (2004). Complete polarization of single
intestinal epithelial cells upon activation of LKB1 by STRAD. Cell 116,
457-466.
Bardeesy, N., Sinha, M., Hezel, A. F., Signoretti, S., Hathaway, N. A.,
Sharpless, N. E., Loda, M., Carrasco, D. R. and DePinho, R. A. (2002).
Loss of the LKB1 tumour suppressor provokes intestinal polyposis but
resistance to transformation. Nature 419, 162-167.
Bateman, A. (1997). The structure of a domain common to archaebacteria and
the homocystinuria disease protein. Trends Biochem. Sci. 22, 12-13.
Bergeron, R., Ren, J. M., Cadman, K. S., Moore, I. K., Perret, P., Pypaert,
M., Young, L. H., Semenkovich, C. F. and Shulman, G. I. (2001). Chronic
activation of AMP kinase results in NRF-1 activation and mitochondrial
biogenesis. Am. J. Physiol. 281, E1340-E1346.
Bolster, D. R., Crozier, S. J., Kimball, S. R. and Jefferson, L. S. (2002). AMPactivated protein kinase suppresses protein synthesis in rat skeletal muscle
through downregulated mTOR signaling. J. Biol. Chem. 277, 23977-23980.
Boudeau, J., Baas, A. F., Deak, M., Morrice, N. A., Kieloch, A.,
Schutowski, M., Prescott, A. R., Clevers, H. C. and Alessi, D. R. (2003).
MO25α/β interact with STRADα/β enhancing their ability to bind, activate
and localize LKB1 in the cytoplasm. EMBO J. 22, 5102-5104.
Browne, G. J., Finn, S. G. and Proud, C. G. (2004). Stimulation of the AMPactivated protein kinase leads to activation of eukaryotic elongation factor
2 kinase and to its phosphorylation at a novel site, serine 398. J. Biol. Chem.
279, 12220-12231.
Brunmair, B., Staniek, K., Gras, F., Scharf, N., Althaym, A., Clara, R.,
Roden, M., Gnaiger, E., Nohl, H., Waldhausl, W. et al. (2004).
Thiazolidinediones, like metformin, inhibit respiratory complex I: a
common mechanism contributing to their antidiabetic actions? Diabetes 53,
1052-1059.
5486
Journal of Cell Science 117 (23)
Buttgereit, F. and Brand, M. D. (1995). A hierarchy of ATP-consuming
processes in mammalian cells. Biochem. J. 312, 163-167.
Carling, D. (2004). The AMP-activated protein kinase cascade – a unifying
system for energy control. Trends Biochem. Sci. 29, 18-24.
Carrera, A. C. (2004). TOR signaling in mammals. J. Cell Sci. 117, 46154616.
Chock, P. B. and Stadtman, E. R. (1977). Superiority of interconvertible
enzyme cascades in metabolite regulation: analysis of multicyclic systems.
Proc. Natl. Acad. Sci. USA 74, 2766-2770.
Choi, S. L., Kim, S. J., Lee, K. T., Kim, J., Mu, J., Birnbaum, M. J., SooKim, S. and Ha, J. (2001). The regulation of AMP-activated protein kinase
by H2O2. Biochem. Biophys. Res. Commun. 287, 92-97.
Corton, J. M., Gillespie, J. G. and Hardie, D. G. (1994). Role of the AMPactivated protein kinase in the cellular stress response. Curr. Biol. 4, 315324.
Crute, B. E., Seefeld, K., Gamble, J., Kemp, B. E. and Witters, L. A.
(1998). Functional domains of the alpha1 catalytic subunit of the AMPactivated protein kinase. J. Biol. Chem. 273, 35347-35354.
Daniel, T. D. and Carling, D. (2002). Functional analysis of mutations in the
γ2 subunit of AMP-activated protein kinase associated with cardiac
hypertrophy and Wolff-Parkinson-White syndrome. J. Biol. Chem. 277,
51017-51024.
Davies, S. P., Helps, N. R., Cohen, P. T. W. and Hardie, D. G. (1995). 5′AMP inhibits dephosphorylation, as well as promoting phosphorylation, of
the AMP-activated protein kinase. Studies using bacterially expressed
human protein phosphatase-2Cα and native bovine protein phosphatase2AC. FEBS Lett. 377, 421-425.
DeVit, M. J. and Johnston, M. (1999). The nuclear exportin Msn5 is required
for nuclear export of the Mig1 glucose repressor of Saccharomyces
cerevisiae. Curr. Biol. 9, 1231-1241.
Estruch, F., Treitel, M. A., Yang, X. and Carlson, M. (1992). N-terminal
mutations modulate yeast SNF1 protein kinase function. Genetics 132, 639650.
Fryer, L. G., Parbu-Patel, A. and Carling, D. (2002). The anti-diabetic drugs
rosiglitazone and metformin stimulate AMP-activated protein kinase
through distinct pathways. J. Biol. Chem. 277, 25226-25232.
Gao, X., Zhang, Y., Arrazola, P., Hino, O., Kobayashi, T., Yeung, R. S.,
Ru, B. and Pan, D. (2002). Tsc tumour suppressor proteins antagonize
amino-acid-TOR signalling. Nat. Cell Biol. 4, 699-704.
Goldbeter, A. and Koshland, D. E. (1981). An amplified sensitivity arising
from covalent modification in biological systems. Proc. Natl. Acad. Sci. USA
78, 6840-6844.
Hardie, D. G. (2003). Minireview: the AMP-activated protein kinase
cascade: the key sensor of cellular energy status. Endocrinology 144,
5179-5183.
Hardie, D. G. (2004a). AMP-activated protein kinase: a key system mediating
metabolic responses to exercise. Med. Sci. Sports Exerc. 36, 28-34.
Hardie, D. G. (2004b). AMP-activated protein kinase: a master switch in
glucose and lipid metabolism. Rev. Endocr. Metab. Disord. 5, 119-125.
Hardie, D. G. and Hawley, S. A. (2001). AMP-activated protein kinase: the
energy charge hypothesis revisited. Bioessays 23, 1112-1119.
Hardie, D. G., Carling, D. and Carlson, M. (1998). The AMPactivated/SNF1 protein kinase subfamily: metabolic sensors of the
eukaryotic cell? Annu. Rev. Biochem. 67, 821-855.
Hardie, D. G., Salt, I. P., Hawley, S. A. and Davies, S. P. (1999). AMPactivated protein kinase: an ultrasensitive system for monitoring cellular
energy charge. Biochem. J. 338, 717-722.
Hardie, D. G., Scott, J. W., Pan, D. A. and Hudson, E. R. (2003).
Management of cellular energy by the AMP-activated protein kinase system.
FEBS Lett. 546, 113-120.
Hawley, S. A., Davison, M., Woods, A., Davies, S. P., Beri, R. K., Carling,
D. and Hardie, D. G. (1996). Characterization of the AMP-activated protein
kinase kinase from rat liver, and identification of threonine-172 as the major
site at which it phosphorylates and activates AMP-activated protein kinase.
J. Biol. Chem. 271, 27879-27887.
Hawley, S. A., Gadalla, A. E., Olsen, G. S. and Hardie, D. G. (2002). The
anti-diabetic drug metformin activates the AMP-activated protein kinase
cascade via an adenine nucleotide-independent mechanism. Diabetes 51,
2420-2425.
Hawley, S. A., Boudeau, J., Reid, J. L., Mustard, K. J., Udd, L., Makela,
T. P., Alessi, D. R. and Hardie, D. G. (2003). Complexes between the
LKB1 tumor suppressor, STRADα/β and MO25α/β are upstream kinases
in the AMP-activated protein kinase cascade. J. Biol. 2, 28.
Hemminki, A., Markie, D., Tomlinson, I., Avizienyte, E., Roth, S., Loukola,
A., Bignell, G., Warren, W., Aminoff, M., Hoglund, P. et al. (1998). A
serine/threonine kinase gene defective in Peutz-Jeghers syndrome. Nature
391, 184-187.
Hong, S. P., Leiper, F. C., Woods, A., Carling, D. and Carlson, M. (2003a).
Activation of yeast Snf1 and mammalian AMP-activated protein kinase by
upstream kinases. Proc. Natl. Acad. Sci. USA 100, 8839-8843.
Hong, Y. H., Varanasi, U. S., Yang, W. and Leff, T. (2003b). AMP-activated
protein kinase regulates HNF4alpha transcriptional activity by inhibiting
dimer formation and decreasing protein stability. J. Biol. Chem. 278, 2749527501.
Horman, S., Browne, G., Krause, U., Patel, J., Vertommen, D., Bertrand,
L., Lavoinne, A., Hue, L., Proud, C. and Rider, M. (2002). Activation
of AMP-activated protein kinase leads to the phosphorylation of
Elongation Factor 2 and an inhibition of protein synthesis. Curr. Biol. 12,
1419-1423.
Hudson, E. R., Pan, D. A., James, J., Lucocq, J. M., Hawley, S. A., Green,
K. A., Baba, O., Terashima, T. and Hardie, D. G. (2003). A novel domain
in AMP-activated protein kinase causes glycogen storage bodies similar to
those seen in hereditary cardiac arrhythmias. Curr. Biol. 13, 861-866.
Hutber, C. A., Hardie, D. G. and Winder, W. W. (1997). Electrical
stimulation inactivates muscle acetyl-CoA carboxylase and increases AMPactivated protein kinase activity. Am. J. Physiol. 272, E262-E266.
Imamura, K., Ogura, T., Kishimoto, A., Kaminishi, M. and Esumi, H.
(2001). Cell cycle regulation via p53 phosphorylation by a 5′-AMP activated
protein kinase activator, 5-aminoimidazole-4-carboxamide-1-beta-dribofuranoside, in a human hepatocellular carcinoma cell line. Biochem.
Biophys. Res. Commun. 287, 562-567.
Inoki, K., Zhu, T. and Guan, K. L. (2003). TSC2 mediates cellular energy
response to control cell growth and survival. Cell 115, 577-590.
Jenne, D. E., Reimann, H., Nezu, J., Friedel, W., Loff, S., Jeschke, R.,
Muller, O., Back, W. and Zimmer, M. (1998). Peutz-Jeghers syndrome is
caused by mutations in a novel serine threonine kinase. Nat. Genet. 18, 3843.
Jiang, R. and Carlson, M. (1997). The Snf1 protein kinase and its activating
subunit, Snf4, interact with distinct domains of the Sip1/Sip2/Gal83
component in the kinase complex. Mol. Cell. Biol. 17, 2099-2106.
Kemp, B. E. (2004). Bateman domains and adenosine derivatives form a
binding contract. J. Clin. Invest. 113, 182-184.
Kimura, N., Tokunaga, C., Dalal, S., Richardson, C., Yoshino, K., Hara,
K., Kemp, B. E., Witters, L. A., Mimura, O. and Yonezawa, K. (2003).
A possible linkage between AMP-activated protein kinase (AMPK) and
mammalian target of rapamycin (mTOR) signalling pathway. Genes Cells
8, 65-79.
Krause, U., Bertrand, L. and Hue, L. (2002). Control of p70 ribosomal
protein S6 kinase and acetyl-CoA carboxylase by AMP-activated protein
kinase and protein phosphatases in isolated hepatocytes. Eur. J. Biochem.
269, 3751-3759.
Kudo, N., Barr, A. J., Barr, R. L., Desai, S. and Lopaschuk, G. D. (1995).
High rates of fatty acid oxidation during reperfusion of ischemic hearts are
associated with a decrease in malonyl-CoA levels due to an increase in 5′AMP-activated protein kinase inhibition of acetyl-CoA carboxylase. J. Biol.
Chem. 270, 17513-17520.
Leclerc, I., Lenzner, C., Gourdon, L., Vaulont, S., Kahn, A. and Viollet,
B. (2001). Hepatocyte nuclear factor-4α involved in type 1 maturity-onset
diabetes of the young is a novel target of AMP-activated protein kinase.
Diabetes 50, 1515-1521.
Lizcano, J. M., Göransson, O., Toth, R., Deak, M., Morrice, N. A.,
Boudeau, J., Hawley, S. A., Udd, L., Mäkelä, T. P., Hardie, D. G. et al.
(2004). LKB1 is a master kinase that activates 13 protein kinases of the
AMPK subfamily, including the MARK/PAR-1 kinases. EMBO J. 23, 833843.
Manning, G., Whyte, D. B., Martinez, R., Hunter, T. and Sudarsanam, S.
(2002). The protein kinase complement of the human genome. Science 298,
1912-1934.
Marsin, A. S., Bertrand, L., Rider, M. H., Deprez, J., Beauloye, C.,
Vincent, M. F., van den Berghe, G., Carling, D. and Hue, L. (2000).
Phosphorylation and activation of heart PFK-2 by AMPK has a role in the
stimulation of glycolysis during ischaemia. Curr. Biol. 10, 1247-1255.
Matejkova, O., Mustard, K. J., Sponarova, J., Flachs, P., Rossmeisl, M.,
Miksik, I., Thomason-Hughes, M., Hardie, D. G. and Kopecky, J.
(2004). Possible involvement of AMP-activated protein kinase in obesity
resistance induced by respiratory uncoupling in white fat. FEBS Lett. 569,
245-248.
Merrill, G. M., Kurth, E., Hardie, D. G. and Winder, W. W. (1997). AICAR
AMP-activated protein kinase
decreases malonyl-CoA and increases fatty acid oxidation in skeletal muscle
of the rat. Am. J. Physiol. 273, E1107-E1112.
Minokoshi, Y., Kim, Y. B., Peroni, O. D., Fryer, L. G., Muller, C., Carling,
D. and Kahn, B. B. (2002). Leptin stimulates fatty-acid oxidation by
activating AMP-activated protein kinase. Nature 415, 339-343.
Minokoshi, Y., Alquier, T., Furukawa, N., Kim, Y. B., Lee, A., Xue, B., Mu,
J., Foufelle, F., Ferre, P., Birnbaum, M. J. et al. (2004). AMP-kinase
regulates food intake by responding to hormonal and nutrient signals in the
hypothalamus. Nature 428, 569-574.
Miyoshi, H., Nakau, M., Ishikawa, T. O., Seldin, M. F., Oshima, M. and
Taketo, M. M. (2002). Gastrointestinal hamartomatous polyposis in LKB1
heterozygous knockout mice. Cancer Res. 62, 2261-2266.
Nakau, M., Miyoshi, H., Seldin, M. F., Imamura, M., Oshima, M. and
Taketo, M. M. (2002). Hepatocellular carcinoma caused by loss of
heterozygosity in LKB1 gene knockout mice. Cancer Res. 62, 4549-4553.
Orci, L., Cook, W. S., Ravazzola, M., Wang, M. Y., Park, B. H.,
Montesano, R. and Unger, R. H. (2004). Rapid transformation of white
adipocytes into fat-oxidizing machines. Proc. Natl. Acad. Sci. USA 101,
2058-2063.
Owen, M. R., Doran, E. and Halestrap, A. P. (2000). Evidence that
metformin exerts its anti-diabetic effects through inhibition of complex 1 of
the mitochondrial respiratory chain. Biochem. J. 348, 607-614.
Pan, D. A. and Hardie, D. G. (2002). A homologue of AMP-activated protein
kinase in Drosophila melanogaster is sensitive to AMP and is activated by
ATP depletion. Biochem. J. 367, 179-186.
Polekhina, G., Gupta, A., Michell, B. J., van Denderen, B., Murthy, S.,
Feil, S. C., Jennings, I. G., Campbell, D. J., Witters, L. A., Parker, M.
W. et al. (2003). AMPK β-subunit targets metabolic stress-sensing to
glycogen. Curr. Biol. 13, 867-871.
Rossi, D. J., Ylikorkala, A., Korsisaari, N., Salovaara, R., Luukko, K.,
Launonen, V., Henkemeyer, M., Ristimaki, A., Aaltonen, L. A. and
Makela, T. P. (2002). Induction of cyclooxygenase-2 in a mouse model of
Peutz-Jeghers polyposis. Proc. Natl. Acad. Sci. USA 99, 12327-12332.
Saha, A. K., Avilucea, P. R., Ye, J. M., Assifi, M. M., Kraegen, E. W. and
Ruderman, N. B. (2004). Pioglitazone treatment activates AMP-activated
protein kinase in rat liver and adipose tissue in vivo. Biochem. Biophys. Res.
Commun. 314, 580-585.
Sakamoto, K., Goransson, O., Hardie, D. G. and Alessi, D. R. (2004).
Activity of LKB1 and AMPK-related kinases in skeletal muscle; effects of
contraction, phenformin and AICAR. Am. J. Physiol. Endocrinol Metab.
287, 310-317.
Salt, I. P., Johnson, G., Ashcroft, S. J. H. and Hardie, D. G. (1998). AMPactivated protein kinase is activated by low glucose in cell lines derived from
pancreatic β cells, and may regulate insulin release. Biochem. J. 335, 533539.
Schrauwen, P., Hardie, D. G., Roorda, B., Clapham, J. C., Abuin, A.,
Thomason-Hughes, M., Green, K., Frederik, P. M. and Hesselink, M.
K. (2004). Improved glucose homeostasis in mice overexpressing human
UCP3: a role for AMP-kinase? Int. J. Obes. Relat. Metab. Disord. 28, 824828.
Scott, J. W., Hawley, S. A., Green, K. A., Anis, M., Stewart, G., Scullion,
G. A., Norman, D. G. and Hardie, D. G. (2004). CBS domains form
energy-sensing modules whose binding of adenosine ligands is disrupted by
disease mutations. J. Clin. Invest. 113, 274-284.
Shaw, R. J., Kosmatka, M., Bardeesy, N., Hurley, R. L., Witters, L. A.,
DePinho, R. A. and Cantley, L. C. (2004). The tumor suppressor LKB1
kinase directly activates AMP-activated kinase and regulates apoptosis in
response to energy stress. Proc. Natl. Acad. Sci. USA 101, 3329-3335.
Smith, F. C., Davies, S. P., Wilson, W. A., Carling, D. and Hardie, D. G.
(1999). The SNF1 kinase complex from Saccharomyces cerevisiae
phosphorylates the repressor protein Mig1p in vitro at four sites within or
near Regulatory Domain 1. FEBS Lett. 453, 219-223.
Sugden, C., Crawford, R. M., Halford, N. G. and Hardie, D. G. (1999).
Regulation of spinach SNF1-related (SnRK1) kinases by protein kinases and
phosphatases is associated with phosphorylation of the T loop and is
regulated by 5′-AMP. Plant J. 19, 433-439.
Sutherland, C. M., Hawley, S. A., McCartney, R. R., Leech, A., Stark, M.
J., Schmidt, M. C. and Hardie, D. G. (2003). Elm1p is one of three
upstream kinases for the Saccharomyces cerevisiae SNF1 complex. Curr.
Biol. 13, 1299-1305.
Terada, S., Goto, M., Kato, M., Kawanaka, K., Shimokawa, T. and Tabata,
5487
I. (2002). Effects of low-intensity prolonged exercise on PGC-1 mRNA
expression in rat epitrochlearis muscle. Biochem. Biophys. Res. Commun.
296, 350-354.
Thelander, M., Olsson, T. and Ronne, H. (2004). Snf1-related protein kinase
1 is needed for growth in a normal day-night light cycle. EMBO J. 23, 19001910.
Tiainen, M., Ylikorkala, A. and Makela, T. P. (1999). Growth suppression
by LKB1 is mediated by a G(1) cell cycle arrest. Proc. Natl. Acad. Sci. USA
96, 9248-9251.
Tiainen, M., Vaahtomeri, K., Ylikorkala, A. and Makela, T. P. (2002).
Growth arrest by the LKB1 tumor suppressor: induction of
p21(WAF1/CIP1). Hum. Mol. Genet. 11, 1497-1504.
Tomas, E., Tsao, T. S., Saha, A. K., Murrey, H. E., Zhang, C. C., Itani, S.
I., Lodish, H. F. and Ruderman, N. B. (2002). Enhanced muscle fat
oxidation and glucose transport by ACRP30 globular domain: acetyl-CoA
carboxylase inhibition and AMP-activated protein kinase activation. Proc.
Natl. Acad. Sci. USA 99, 16309-16313.
Vavvas, D., Apazidis, A., Saha, A. K., Gamble, J., Patel, A., Kemp, B. E.,
Witters, L. A. and Ruderman, N. B. (1997). Contraction-induced changes
in acetyl-CoA carboxylase and 5′-AMP-activated kinase in skeletal muscle.
J. Biol. Chem. 272, 13255-13261.
Wang, W., Fan, J., Yang, X., Fürer, S., de Silanes, I. L., von Kobbe, C.,
Guo, J., Georas, S., Foufelle, F., Hardie, D. G. et al. (2002). AMPactivated kinase regulates HuR subcellular localization. Mol. Cell. Biol. 27,
3425-3436.
Wang, W., Yang, X., Lopez de Silanes, I., Carling, D. and Gorospe, M.
(2003). Increased AMP:ATP ratio and AMP-activated protein kinase activity
during cellular senescence linked to reduced HuR function. J. Biol. Chem.
278, 27016-27023.
Winder, W. W. and Hardie, D. G. (1996). Inactivation of acetyl-CoA
carboxylase and activation of AMP-activated protein kinase in muscle
during exercise. Am. J. Physiol. 270, E299-E304.
Witters, L. A., Nordlund, A. C. and Marshall, L. (1991). Regulation of
intracellular acetyl-CoA carboxylase by ATP depletors mimics the action of
the 5′-AMP-activated protein kinase. Biochem. Biophys. Res. Commun. 181,
1486-1492.
Wojtaszewski, J. F. P., Jørgensen, S. B., Hellsten, Y., Hardie, D. G. and
Richter, E. A. (2002). Glycogen-dependent effects of AICA riboside on
AMP-activated protein kinase and glycogen synthase activities in rat skeletal
muscle. Diabetes 51, 284-292.
Wojtaszewski, J. F., MacDonald, C., Nielsen, J. N., Hellsten, Y., Hardie,
D. G., Kemp, B. E., Kiens, B. and Richter, E. A. (2003). Regulation of
5′AMP-activated protein kinase activity and substrate utilization in
exercising human skeletal muscle. Am. J. Physiol. 284, E813-E822.
Woods, A., Johnstone, S. R., Dickerson, K., Leiper, F. C., Fryer, L. G.,
Neumann, D., Schlattner, U., Wallimann, T., Carlson, M. and Carling,
D. (2003). LKB1 is the upstream kinase in the AMP-activated protein kinase
cascade. Curr. Biol. 13, 2004-2008.
Yamauchi, T., Kamon, J., Minokoshi, Y., Ito, Y., Waki, H., Uchida, S.,
Yamashita, S., Noda, M., Kita, S., Ueki, K. et al. (2002). Adiponectin
stimulates glucose utilization and fatty-acid oxidation by activating AMPactivated protein kinase. Nat. Med. 6, 1288-1295.
Yang, W., Hong, Y. H., Shen, X. Q., Frankowski, C., Camp, H. S. and Leff,
T. (2001). Regulation of transcription by AMP-activated protein kinase.
Phosphorylation of p300 blocks its interaction with nuclear receptors. J.
Biol. Chem. 276, 38341-38344.
Zhang, R., Evans, G., Rotella, F. J., Westbrook, E. M., Beno, D.,
Huberman, E., Joachimiak, A. and Collart, F. R. (1999). Characteristics
and crystal structure of bacterial inosine-5′-monophosphate dehydrogenase.
Biochemistry 38, 4691-4700.
Zheng, D., MacLean, P. S., Pohnert, S. C., Knight, J. B., Olson, A. L.,
Winder, W. W. and Dohm, G. L. (2001). Regulation of muscle GLUT-4
transcription by AMP-activated protein kinase. J. Appl. Physiol. 91, 10731083.
Zhou, G., Myers, R., Li, Y., Chen, Y., Shen, X., Fenyk-Melody, J., Wu, M.,
Ventre, J., Doebber, T., Fujii, N. et al. (2001). Role of AMP-activated protein
kinase in mechanism of metformin action. J. Clin. Invest. 108, 1167-1174.
Zong, H., Ren, J. M., Young, L. H., Pypaert, M., Mu, J., Birnbaum, M. J.
and Shulman, G. I. (2002). AMP kinase is required for mitochondrial
biogenesis in skeletal muscle in response to chronic energy deprivation.
Proc. Natl. Acad. Sci. USA 99, 15983-15987.